首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
2.
Bats are hosts to a variety of viruses capable of zoonotic transmissions. Because of increased contact between bats, humans, and other animal species, the possibility exists for further cross-species transmissions and ensuing disease outbreaks. We describe here full and partial viral genomes identified using metagenomics in the guano of bats from California and Texas. A total of 34% and 58% of 390,000 sequence reads from bat guano in California and Texas, respectively, were related to eukaryotic viruses, and the largest proportion of those infect insects, reflecting the diet of these insectivorous bats, including members of the viral families Dicistroviridae, Iflaviridae, Tetraviridae, and Nodaviridae and the subfamily Densovirinae. The second largest proportion of virus-related sequences infects plants and fungi, likely reflecting the diet of ingested insects, including members of the viral families Luteoviridae, Secoviridae, Tymoviridae, and Partitiviridae and the genus Sobemovirus. Bat guano viruses related to those infecting mammals comprised the third largest group, including members of the viral families Parvoviridae, Circoviridae, Picornaviridae, Adenoviridae, Poxviridae, Astroviridae, and Coronaviridae. No close relative of known human viral pathogens was identified in these bat populations. Phylogenetic analysis was used to clarify the relationship to known viral taxa of novel sequences detected in bat guano samples, showing that some guano viral sequences fall outside existing taxonomic groups. This initial characterization of the bat guano virome, the first metagenomic analysis of viruses in wild mammals using second-generation sequencing, therefore showed the presence of previously unidentified viral species, genera, and possibly families. Viral metagenomics is a useful tool for genetically characterizing viruses present in animals with the known capability of direct or indirect viral zoonosis to humans.Bats belong to one of the most diverse, abundant, and widely distributed group of mammals. More than 1,100 bat species belong to the order of Chiroptera, representing approximately 20% of all mammalian species (54). Most bat species feed on insects and other arthropods, while others feed on fruit nectar, bird or mammal blood, and small vertebrates such as fish, frogs, mice, and birds (30). Of the 47 species of bats reported in the United States, most of them are insectivorous (http://www.batcon.org/).Bats are considered the natural reservoir of a large variety of zoonotic viruses causing serious human diseases such as lyssaviruses, henipaviruses, severe acute respiratory syndrome coronavirus, and Ebola virus (6, 38, 46, 59, 63, 65). Characteristics of bats, including their genetic diversity, broad geological distribution, gregarious habits, high population density, migratory habits, and long life span (30, 58), likely endow them with the ability to host diverse viruses, some of which are also able to infect humans and other mammals (41, 63).More than 80 virus species have been isolated or detected in bats using nucleic acid-based methods (6, 38, 59, 65). Viruses that have been recently discovered in bats include astroviruses, adeno-associated viruses (AAVs), adenoviruses, herpesviruses, and polyomavirus (8, 9, 13, 31, 32, 35, 37, 39, 40, 42, 61, 62, 68). For example, it was recently reported that a newly identified adenovirus isolated from bat guano was capable of infecting various vertebrate cell lines, including those of humans, monkeys, dogs, and pigs (35). With increasing human populations in previously wild areas, contact of bats with humans and with wild and domestic animals has increased, providing greater opportunities for cross-species transmissions of potentially pathogenic bat viruses. To better understand the range of viruses carried by bats, we undertook an initial characterization of the guano viromes of several common bat species in the United States.The development of massively parallel sequencing technology makes is possible to reveal uncultured viral assemblages within biological or environmental samples (11, 28). To date, this approach has been used to characterize viruses in equine feces (7), human blood (5), tissue (14), human feces (3, 4, 15, 45, 60, 67), and human respiratory secretions (64), which in turn has facilitated the discovery of many novel viruses (18, 20, 25, 33, 47, 50). In the present study, we analyzed the viruses present in guano from several bat species in California and Texas, using sequence-independent PCR amplification, pyrosequencing, and sequence similarity searches.  相似文献   

3.
Avian H7 influenza viruses have been responsible for poultry outbreaks worldwide and have resulted in numerous cases of human infection in recent years. The high rate of conjunctivitis associated with avian H7 subtype virus infections may represent a portal of entry for avian influenza viruses and highlights the need to better understand the apparent ocular tropism observed in humans. To study this, mice were inoculated by the ocular route with viruses of multiple subtypes and degrees of virulence. We found that in contrast to human (H3N2 and H1N1) viruses, H7N7 viruses isolated from The Netherlands in 2003 and H7N3 viruses isolated from British Columbia, Canada, in 2004, two subtypes that were highly virulent for poultry, replicated to a significant titer in the mouse eye. Remarkably, an H7N7 virus, as well as some avian H5N1 viruses, spread systemically following ocular inoculation, including to the brain, resulting in morbidity and mortality of mice. This correlated with efficient replication of highly pathogenic H7 and H5 subtypes in murine corneal epithelial sheets (ex vivo) and primary human corneal epithelial cells (in vitro). Influenza viruses were labeled to identify the virus attachment site in the mouse cornea. Although we found abundant H7 virus attachment to corneal epithelial tissue, this did not account for the differences in virus replication as multiple subtypes were able to attach to these cells. These findings demonstrate that avian influenza viruses within H7 and H5 subtypes are capable of using the eye as a portal of entry.Highly pathogenic avian influenza (HPAI) H5N1 viruses, which have resulted in over 420 documented cases of human infection to date, have generally caused acute, often severe and fatal, respiratory illness (1, 50). While conjunctivitis following infection with H5N1 or human influenza viruses has been rare, most human infections associated with H7 subtype viruses have resulted in ocular and not respiratory disease (1, 9, 37, 38). Infrequent reports of human conjunctivitis infection following exposure to H7 influenza viruses date from 1977, predominantly resulting from laboratory or occupational exposure (21, 40, 48). However, in The Netherlands in 2003, more than 80 human infections with H7N7 influenza virus occurred among poultry farmers and cullers amid widespread outbreaks of HPAI in domestic poultry; the majority of these human infections resulted in conjunctivitis (14, 20). Additionally, conjunctivitis was documented in the two human infections resulting from an H7N3 outbreak in British Columbia, Canada, in 2004, as well as in H7N3- and H7N2-infected individuals in the United Kingdom in 2006 and 2007, respectively (13, 18, 29, 46, 51). The properties that contribute to an apparent ocular tropism of some influenza viruses are currently not well understood (30).Host cell glycoproteins bearing sialic acids (SAs) are the cellular receptors for influenza viruses and can be found on epithelial cells within both the human respiratory tract and ocular tissue (26, 31, 41). Both respiratory and ocular tissues additionally secrete sialylated mucins that function in pathogen defense and protection of the epithelial surface (5, 11, 22). Within the upper respiratory tract, α2-6-linked SAs (the preferred receptor for human influenza viruses) predominate on epithelial cells (26). While α2-3-linked SAs are also present to a lesser degree on respiratory epithelial cells, this linkage is more abundantly expressed on secreted mucins (2). In contrast, α2-3-linked SAs (the preferred receptor for avian influenza viruses) are found on corneal and conjunctival epithelial cells of the human eye (31, 41), while secreted ocular mucins are abundantly composed of α2-6 SAs (5). It has been suggested that avian influenza viruses are more suited to infect the ocular surface due to their general α2-3-linked SA binding preference, but this has not been demonstrated experimentally (30).The mouse model has been used previously to study the role of ocular exposure to respiratory viruses (6, 39). In mice, ocular inoculation with an H3N2 influenza virus resulted in virus replication in nasal turbinates and lung (39), whereas ocular infection with respiratory syncytial virus (RSV) resulted in detectable virus titers in the eye and lung (6). These studies have revealed that respiratory viruses are not limited to the ocular area following inoculation at this site. However, the ability of influenza viruses to replicate specifically within ocular tissue has not been examined.Despite repeated instances of conjunctivitis associated with H7 subtype infections in humans, the reasons for this apparent ocular tropism have not been studied extensively. Here, we present a murine model to study the ability of human and avian influenza viruses to cause disease by the ocular route. We found that highly pathogenic H7 and H5 influenza viruses were capable of causing a systemic and lethal infection in mice following ocular inoculation. These highly pathogenic viruses, unlike human H3N2 and H1N1 viruses, replicated to significant titers in the mouse corneal epithelium and primary human corneal epithelial cells (HCEpiCs). Identification of viruses well suited to infecting the ocular surface is the first step in better understanding the ability of influenza viruses of multiple subtypes to use this tissue as a portal of entry.  相似文献   

4.
In their vertebrate hosts, arboviruses such as Semliki Forest virus (SFV) (Togaviridae) generally counteract innate defenses and trigger cell death. In contrast, in mosquito cells, following an early phase of efficient virus production, a persistent infection with low levels of virus production is established. Whether arboviruses counteract RNA interference (RNAi), which provides an important antiviral defense system in mosquitoes, is an important question. Here we show that in Aedes albopictus-derived mosquito cells, SFV cannot prevent the establishment of an antiviral RNAi response or prevent the spread of protective antiviral double-stranded RNA/small interfering RNA (siRNA) from cell to cell, which can inhibit the replication of incoming virus. The expression of tombusvirus siRNA-binding protein p19 by SFV strongly enhanced virus spread between cultured cells rather than virus replication in initially infected cells. Our results indicate that the spread of the RNAi signal contributes to limiting virus dissemination.In animals, RNA interference (RNAi) was first described for Caenorhabditis elegans (27). The production or introduction of double-stranded RNA (dsRNA) in cells leads to the degradation of mRNAs containing homologous sequences by sequence-specific cleavage of mRNAs. Central to RNAi is the production of 21- to 26-nucleotide small interfering RNAs (siRNAs) from dsRNA and the assembly of an RNA-induced silencing complex (RISC), followed by the degradation of the target mRNA (23, 84). RNAi is a known antiviral strategy of plants (3, 53) and insects (21, 39, 51). Study of Drosophila melanogaster in particular has given important insights into RNAi responses against pathogenic viruses and viral RNAi inhibitors (31, 54, 83, 86, 91). RNAi is well characterized for Drosophila, and orthologs of antiviral RNAi genes have been found in Aedes and Culex spp. (13, 63).Arboviruses, or arthropod-borne viruses, are RNA viruses mainly of the families Bunyaviridae, Flaviviridae, and Togaviridae. The genus Alphavirus within the family Togaviridae contains several mosquito-borne pathogens: arboviruses such as Chikungunya virus (16) and equine encephalitis viruses (88). Replication of the prototype Sindbis virus and Semliki Forest virus (SFV) is well understood (44, 71, 74, 79). Their genome consists of a positive-stranded RNA with a 5′ cap and a 3′ poly(A) tail. The 5′ two-thirds encodes the nonstructural polyprotein P1234, which is cleaved into four replicase proteins, nsP1 to nsP4 (47, 58, 60). The structural polyprotein is encoded in the 3′ one-third of the genome and cleaved into capsid and glycoproteins after translation from a subgenomic mRNA (79). Cytoplasmic replication complexes are associated with cellular membranes (71). Viruses mature by budding at the plasma membrane (35).In nature, arboviruses are spread by arthropod vectors (predominantly mosquitoes, ticks, flies, and midges) to vertebrate hosts (87). Little is known about how arthropod cells react to arbovirus infection. In mosquito cell cultures, an acute phase with efficient virus production is generally followed by the establishment of a persistent infection with low levels of virus production (9). This is fundamentally different from the cytolytic events following arbovirus interactions with mammalian cells and pathogenic insect viruses with insect cells. Alphaviruses encode host response antagonists for mammalian cells (2, 7, 34, 38).RNAi has been described for mosquitoes (56) and, when induced before infection, antagonizes arboviruses and their replicons (1, 4, 14, 15, 29, 30, 32, 42, 64, 65). RNAi is also functional in various mosquito cell lines (1, 8, 43, 49, 52). In the absence of RNAi, alphavirus and flavivirus replication and/or dissemination is enhanced in both mosquitoes and Drosophila (14, 17, 31, 45, 72). RNAi inhibitors weakly enhance SFV replicon replication in tick and mosquito cells (5, 33), posing the questions of how, when, and where RNAi interferes with alphavirus infection in mosquito cells.Here we use an A. albopictus-derived mosquito cell line to study RNAi responses to SFV. Using reporter-based assays, we demonstrate that SFV cannot avoid or efficiently inhibit the establishment of an RNAi response. We also demonstrate that the RNAi signal can spread between mosquito cells. SFV cannot inhibit cell-to-cell spread of the RNAi signal, and spread of the virus-induced RNAi signal (dsRNA/siRNA) can inhibit the replication of incoming SFV in neighboring cells. Furthermore, we show that SFV expression of a siRNA-binding protein increases levels of virus replication mainly by enhancing virus spread between cells rather than replication in initially infected cells. Taken together, these findings suggest a novel mechanism, cell-to-cell spread of antiviral dsRNA/siRNA, by which RNAi limits SFV dissemination in mosquito cells.  相似文献   

5.
6.
Cytosolic chaperones are a diverse group of ubiquitous proteins that play central roles in multiple processes within the cell, including protein translation, folding, intracellular trafficking, and quality control. These cellular proteins have also been implicated in the replication of numerous viruses, although the full extent of their involvement in viral replication is unknown. We have previously shown that the heat shock protein 40 (hsp40) chaperone encoded by the yeast YDJ1 gene facilitates RNA replication of flock house virus (FHV), a well-studied and versatile positive-sense RNA model virus. To further explore the roles of chaperones in FHV replication, we examined a panel of 30 yeast strains with single deletions of cytosolic proteins that have known or hypothesized chaperone activity. We found that the majority of cytosolic chaperone deletions had no impact on FHV RNA accumulation, with the notable exception of J-domain-containing hsp40 chaperones, where deletion of APJ1 reduced FHV RNA accumulation by 60%, while deletion of ZUO1, JJJ1, or JJJ2 markedly increased FHV RNA accumulation, by 4- to 40-fold. Further studies using cross complementation and double-deletion strains revealed that the contrasting effects of J domain proteins were reproduced by altering expression of the major cytosolic hsp70s encoded by the SSA and SSB families and were mediated in part by divergent effects on FHV RNA polymerase synthesis. These results identify hsp70 chaperones as critical regulators of FHV RNA replication and indicate that cellular chaperones can have both positive and negative regulatory effects on virus replication.The compact genomes of viruses relative to those of other infectious agents restrict their ability to encode all proteins required to complete their replication cycles. To circumvent this limitation, viruses often utilize cellular factors or processes to complete essential steps in replication. One group of cellular proteins frequently targeted by viruses are cellular chaperones, which include a diverse set of heat shock proteins (hsps) that normally facilitate cellular protein translation, folding, trafficking, and degradation (18, 64). The connection between viruses and cellular chaperones was originally identified in bacteria, where the Escherichia coli hsp40 and hsp70 homologues, encoded by dnaJ and dnaK, respectively, were identified as bacterial genes essential for bacteriophage λ DNA replication (62). Research over the past 30 years has further revealed the importance of cellular chaperones in viral replication, such that the list of virus-hsp connections is now quite extensive and includes viruses from numerous families with diverse genome structures (4, 6, 7, 16, 19, 20, 23, 25, 40, 41, 44, 51, 54, 60). These studies have demonstrated the importance of cellular chaperones in multiple steps of the viral life cycle, including entry, viral protein translation, genome replication, encapsidation, and virion release. However, the list of virus-hsp connections is likely incomplete. Further studies to explore this particular host-pathogen interaction will shed light on virus replication mechanisms and pathogenesis, and potentially highlight targets for novel antiviral agents.To study the role of cellular chaperones in the genome replication of positive-sense RNA viruses, we use flock house virus (FHV), a natural insect pathogen and well-studied member of the Nodaviridae family. The FHV life cycle shares many common features with other positive-sense RNA viruses, including the membrane-specific targeting and assembly of functional RNA replication complexes (37, 38), the exploitation of various cellular processes and host factors for viral replication (5, 23, 60), and the induction of large-scale membrane rearrangements (24, 28, 38, 39). FHV virions contain a copackaged bipartite genome consisting of RNA1 (3.1 kb) and RNA2 (1.4 kb), which encode protein A, the viral RNA-dependent RNA polymerase, and the structural capsid protein precursor, respectively (1). During active genome replication, FHV produces a subgenomic RNA3 (0.4 kb), which encodes the RNA interference inhibitor protein B2 (12, 29, 32). These viral characteristics make FHV an excellent model system to study many aspects of positive-sense RNA virus biology.In addition to the benefits of a simple genome, FHV is able to establish robust RNA replication in a wide variety of genetically tractable eukaryotic hosts, including Drosophila melanogaster (38), Caenorhabditis elegans (32), and Saccharomyces cerevisiae (46). The budding yeast S. cerevisiae has been an exceptionally useful model host to study the mechanisms of viral RNA replication complex assembly and function with FHV (31, 37, 39, 45, 53, 55, 56, 60) as well as other positive-sense RNA viruses (11). The facile genetics of S. cerevisiae, along with the vast array of well-defined cellular and molecular tools and techniques, make it an ideal eukaryotic host for the identification of cellular factors required for positive-sense RNA virus replication. Furthermore, readily available yeast libraries with deletions and regulated expression of individual proteins have led to the completion of several high-throughput screens to provide a global survey of host factors that impact virus replication (26, 42, 52). An alternative approach with these yeast libraries that reduces the inherently high false-negative rates associated with high-throughput screens is to focus on a select set of host genes associated with a particular cellular pathway, process, or location previously implicated in virus replication.We have utilized such a targeted approach and focused on examining the impact of cytosolic chaperones on FHV RNA replication. Previously, we have shown that the cellular chaperone hsp90 facilitates protein A synthesis in Drosophila cells (5, 23), and the hsp40 encoded by the yeast YDJ1 gene facilitates FHV RNA replication in yeast, in part through effects on both protein A accumulation and function (60). In this report, we further extend these observations by examining FHV RNA accumulation in a panel of yeast strains with deletions of known or hypothesized cytosolic chaperones. We demonstrate that cytosolic chaperones can have either suppressive or enhancing effects on FHV RNA accumulation. In particular, related hsp70 members encoded by the SSA and SSB yeast chaperone families have marked and dramatically divergent effects on both genomic and subgenomic RNA accumulation and viral polymerase synthesis. These results highlight the complexities of the host-pathogen interactions that influence positive-sense RNA virus replication and identify the hsp70 family of cytosolic chaperones as key regulators of FHV replication.  相似文献   

7.
8.
The highly pathogenic H5N1 avian influenza virus emerged from China in 1996 and has spread across Eurasia and Africa, with a continuous stream of new cases of human infection appearing since the first large-scale outbreak among migratory birds at Qinghai Lake. The role of wild birds, which are the natural reservoirs for the virus, in the epidemiology of the H5N1 virus has raised great public health concern, but their role in the spread of the virus within the natural ecosystem of free-ranging terrestrial wild mammals remains unclear. In this study, we investigated H5N1 virus infection in wild pikas in an attempt to trace the circulation of the virus. Seroepidemiological surveys confirmed a natural H5N1 virus infection of wild pikas in their native environment. The hemagglutination gene of the H5N1 virus isolated from pikas reveals two distinct evolutionary clades, a mixed/Vietnam H5N1 virus sublineage (MV-like pika virus) and a wild bird Qinghai (QH)-like H5N1 virus sublineage (QH-like pika virus). The amino acid residue (glutamic acid) at position 627 encoded by the PB2 gene of the MV-like pika virus was different from that of the QH-like pika virus; the residue of the MV-like pika virus was the same as that of the goose H5N1 virus (A/GS/Guangdong [GD]/1/96). Further, we discovered that in contrast to the MV-like pika virus, which is nonpathogenic to mice, the QH-like pika virus is highly pathogenic. To mimic the virus infection of pikas, we intranasally inoculated rabbits, a species closely related to pikas, with the H5N1 virus of pika origin. Our findings first demonstrate that wild pikas are mammalian hosts exposed to H5N1 subtype avian influenza viruses in the natural ecosystem and also imply a potential transmission of highly pathogenic avian influenza virus from wild mammals into domestic mammalian hosts and humans.Highly pathogenic avian influenza (HPAI) is an extremely infectious, systemic viral disease that causes a high rate of mortality in birds. HPAI H5N1 viruses are now endemic in avian populations in Southeast Asia and have repeatedly been transmitted to humans (9, 14, 27). Since 2003, the H5N1 subtype has been reported in 391 human cases of influenza and has caused 247 human deaths in 15 countries, leading to greater than 60% mortality among infected individuals (38). Although currently incapable of sustained human-to-human transmission, H5N1 viruses undoubtedly pose a serious threat to public health, as well as to the global economy. Hence, preparedness for such a threat is a global priority (36).Wild birds are considered to be natural reservoirs for influenza A viruses (6, 18, 21, 35, 37). Of the 144 type A influenza virus hemagglutinin-neuraminidase (HA-NA) combinations, 103 have been found in wild birds (5, 7, 17, 37). Since the first HPAI outbreak among migratory wild birds appeared at Qinghai Lake in western China in May 2005 (3, 16, 25, 34, 41), HPAI viruses of the H5N1 subtype have been isolated from poultry throughout Eurasia and Africa. The continued occurrence of human cases has created a situation that could facilitate a pandemic emergence. There is heightened concern that wild birds are a reservoir for influenza A viruses that switch hosts and stably adapt to mammals, including horses, swine, and humans (11, 19, 22, 37).Despite the recent expansion of avian influenza virus (AIV) surveillance and genomic data (5, 17, 20, 21, 33, 40), fundamental questions remain concerning the ecology and evolution of these viruses. Little is known about how terrestrial wild mammals within their natural ecological systems affect HPAI H5N1 epidemiology or about the virus''s effects on public health, though there are many reports of natural and experimental H5N1 virus infection in animals belonging to the taxonomic orders Carnivora (12, 13, 15, 28, 29) and Artiodactyla (15). Herein, we provide the results of our investigation into H5N1 virus infection in wild pikas (Ochotona curzoniae of the order Lagomorpha) within their natural ecological setting. We describe our attempt to trace the circulation of H5N1 viruses and to characterize pika H5N1 influenza virus (PK virus).  相似文献   

9.
Mature glycoprotein spikes are inserted in the Lassa virus envelope and consist of the distal subunit GP-1, the transmembrane-spanning subunit GP-2, and the signal peptide, which originate from the precursor glycoprotein pre-GP-C by proteolytic processing. In this study, we analyzed the oligomeric structure of the viral surface glycoprotein. Chemical cross-linking studies of mature glycoprotein spikes from purified virus revealed the formation of trimers. Interestingly, sucrose density gradient analysis of cellularly expressed glycoprotein showed that in contrast to trimeric mature glycoprotein complexes, the noncleaved glycoprotein forms monomers and oligomers spanning a wide size range, indicating that maturation cleavage of GP by the cellular subtilase SKI-1/S1P is critical for formation of the correct oligomeric state. To shed light on a potential relation between cholesterol and GP trimer stability, we performed cholesterol depletion experiments. Although depletion of cholesterol had no effect on trimerization of the glycoprotein spike complex, our studies revealed that the cholesterol content of the viral envelope is important for the infectivity of Lassa virus. Analyses of the distribution of viral proteins in cholesterol-rich detergent-resistant membrane areas showed that Lassa virus buds from membrane areas other than those responsible for impaired infectivity due to cholesterol depletion of lipid rafts. Thus, derivation of the viral envelope from cholesterol-rich membrane areas is not a prerequisite for the impact of cholesterol on virus infectivity.Lassa virus (LASV) is a member of the family Arenaviridae, of which Lymphocytic choriomeningitis virus (LCMV) is the prototype. Arenaviruses comprise more than 20 species, divided into the Old World and New World virus complexes (19). The Old World arenaviruses include the human pathogenic LASV strains, Lujo virus, which was first identified in late 2008 and is associated with an unprecedented high case fatality rate in humans, the nonhuman pathogenic Ippy, Mobala, and Mopeia viruses, and the recently described Kodoko virus (10, 30, 49). The New World virus complex contains, among others, the South American hemorrhagic fever-causing viruses Junín virus, Machupo virus, Guanarito virus, Sabiá virus, and the recently discovered Chapare virus (22).Arenaviruses contain a bisegmented single-stranded RNA genome encoding the polymerase L, matrix protein Z, nucleoprotein NP, and glycoprotein GP. The bipartite ribonucleoprotein of LASV is surrounded by a lipid envelope derived from the plasma membrane of the host cell. The matrix protein Z has been identified as a major budding factor, which lines the interior of the viral lipid membrane, in which GP spikes are inserted (61, 75). The glycoprotein is synthesized as precursor protein pre-GP-C and is cotranslationally cleaved by signal peptidase into GP-C and the signal peptide, which exhibits unusual length, stability, and topology (3, 27, 28, 33, 70, 87). Moreover, the arenaviral signal peptide functions as trans-acting maturation factor (2, 26, 33). After processing by signal peptidase, GP-C of both New World and Old World arenaviruses is cleaved by the cellular subtilase subtilisin kexin isozyme-1/site-1 protease (SKI-1/S1P) into the distal subunit GP-1 and the membrane-anchored subunit GP-2 within the secretory pathway (5, 52, 63). For LCMV, it has been shown that GP-1 subunits are linked to each other by disulfide bonds and are noncovalently connected to GP-2 subunits (14, 24, 31). GP-1 is responsible for binding to the host cell receptor, while GP-2 mediates fusion between the virus envelope and the endosomal membrane at low pH due to a bipartite fusion peptide near the amino terminus (24, 36, 44). Sequence analysis of the LCMV GP-2 ectodomain revealed two heptad repeats that most likely form amphipathic helices important for this process (34, 86).In general, viral class I fusion proteins have triplets of α-helical structures in common, which contain heptad repeats (47, 73). In contrast, class II fusion proteins are characterized by β-sheets that form dimers in the prefusion status and trimers in the postfusion status (43). The class III fusion proteins are trimers that, unlike class I fusion proteins, were not proteolytically processed N-terminally of the fusion peptide, resulting in a fusion-active membrane-anchored subunit (39, 62). Previous studies with LCMV described a tetrameric organization of the glycoprotein spikes (14), while more recent data using a bacterially expressed truncated ectodomain of the LCMV GP-2 subunit pointed toward a trimeric spike structure (31). Due to these conflicting data regarding the oligomerization status of LCMV GP, it remains unclear to which class of fusion proteins the arenaviral glycoproteins belong.The state of oligomerization and the correct conformation of viral glycoproteins are crucial for membrane fusion during virus entry. The early steps of infection have been shown for several viruses to be dependent on the cholesterol content of the participating membranes (i.e., either the virus envelope or the host cell membrane) (4, 9, 15, 20, 21, 23, 40, 42, 53, 56, 76, 78, 79). In fact, it has been shown previously that entry of both LASV and LCMV is susceptible to cholesterol depletion of the target host cell membrane using methyl-β-cyclodextrin (MβCD) treatment (64, 71). Moreover, cholesterol not only plays an important role in the early steps during entry in the viral life cycle but also is critical in the virus assembly and release process. Several viruses of various families, including influenza virus, human immunodeficiency virus type 1 (HIV-1), measles virus, and Ebola virus, use the ordered environment of lipid raft microdomains. Due to their high levels of glycosphingolipids and cholesterol, these domains are characterized by insolubility in nonionic detergents under cold conditions (60, 72). Recent observations have suggested that budding of the New World arenavirus Junin virus occurs from detergent-soluble membrane areas (1). Assembly and release from distinct membrane microdomains that are detergent soluble have also been described for vesicular stomatitis virus (VSV) (12, 38, 68). At present, however, it is not known whether LASV requires cholesterol in its viral envelope for successful virus entry or whether specific membrane microdomains are important for LASV assembly and release.In this study, we first investigated the oligomeric state of the premature and mature LASV glycoprotein complexes. Since it has been shown for several membrane proteins that the oligomerization and conformation are dependent on cholesterol (58, 59, 76, 78), we further analyzed the dependence of the cholesterol content of the virus envelope on glycoprotein oligomerization and virus infectivity. Finally, we characterized the lipid membrane areas from which LASV is released.  相似文献   

10.
Like other RNA viruses, coxsackievirus B5 (CVB5) exists as circulating heterogeneous populations of genetic variants. In this study, we present the reconstruction and characterization of a probable ancestral virion of CVB5. Phylogenetic analyses based on capsid protein-encoding regions (the VP1 gene of 41 clinical isolates and the entire P1 region of eight clinical isolates) of CVB5 revealed two major cocirculating lineages. Ancestral capsid sequences were inferred from sequences of these contemporary CVB5 isolates by using maximum likelihood methods. By using Bayesian phylodynamic analysis, the inferred VP1 ancestral sequence dated back to 1854 (1807 to 1898). In order to study the properties of the putative ancestral capsid, the entire ancestral P1 sequence was synthesized de novo and inserted into the replicative backbone of an infectious CVB5 cDNA clone. Characterization of the recombinant virus in cell culture showed that fully functional infectious virus particles were assembled and that these viruses displayed properties similar to those of modern isolates in terms of receptor preferences, plaque phenotypes, growth characteristics, and cell tropism. This is the first report describing the resurrection and characterization of a picornavirus with a putative ancestral capsid. Our approach, including a phylogenetics-based reconstruction of viral predecessors, could serve as a starting point for experimental studies of viral evolution and might also provide an alternative strategy for the development of vaccines.The group B coxsackieviruses (CVBs) (serotypes 1 to 6) were discovered in the 1950s in a search for new poliovirus-like viruses (33, 61). Infections caused by CVBs are often asymptomatic but may occasionally result in severe diseases of the heart, pancreas, and central nervous system (99). CVBs are small icosahedral RNA viruses belonging to the Human enterovirus B (HEV-B) species within the family Picornaviridae (89). In the positive single-stranded RNA genome, the capsid proteins VP1 to VP4 are encoded within the P1 region, whereas the nonstructural proteins required for virus replication are encoded within the P2 and P3 regions (4). The 30-nm capsid has an icosahedral symmetry and consists of 60 copies of each of the four structural proteins. The VP1, VP2, and VP3 proteins are surface exposed, whereas the VP4 protein lines the interior of the virus capsid (82). The coxsackievirus and adenovirus receptor (CAR), a cell adhesion molecule of the immunoglobulin superfamily, serves as the major cell surface attachment molecule for all six serotypes of CVB (5, 6, 39, 60, 98). Some strains of CVB1, CVB3 and CVB5 also interact with the decay-accelerating factor (DAF) (CD55), a member of the family of proteins that regulate the complement cascade. However, the attachment of CVBs to DAF alone does not permit the infection of cells (6, 7, 59, 85).Picornaviruses exist as genetically highly diverse populations within their hosts, referred to as quasispecies (20, 57). This genetic plasticity enables these viruses to adapt rapidly to new environments, but at the same time, it may compromise the structural integrity and enzymatic functionality of the virus. The selective constraints imposed on the picornavirus genome are reflected in the different regions used for different types of evolutionary studies. The highly conserved RNA-dependent RNA polymerase (3Dpol) gene is used to establish phylogenetic relationships between more-distantly related viruses (e.g., viruses belonging to different genera) (38), whereas the variable genomic sequence encoding the VP1 protein is used for the classification of serotypes (13, 14, 69, 71, 72).In 1963, Pauling and Zuckerkandl proposed that comparative analyses of contemporary protein sequences can be used to predict the sequences of their ancient predecessors (73). Experimental reconstruction of ancestral character states has been applied to evolutionary studies of several different proteins, e.g., galectins (49), G protein-coupled receptors (52), alcohol dehydrogenases (95), rhodopsins (15), ribonucleases (46, 88, 110), elongation factors (32), steroid receptors (10, 96, 97), and transposons (1, 45, 87). In the field of virology, reconstructed ancestral or consensus protein sequences have been used in attempts to develop vaccine candidates for human immunodeficiency virus type 1 (21, 51, 66, 81) but rarely to examine general phenotypic properties.In this study, a CVB5 virus with a probable ancestral virion (CVB5-P1anc) was constructed and characterized. We first analyzed in detail the evolutionary relationships between structural genes of modern CVB5 isolates and inferred a time scale for their evolutionary history. An ancestral virion sequence was subsequently inferred by using a maximum likelihood (ML) method. This sequence was then synthesized de novo, cloned into a replicative backbone of an infectious CVB5 cDNA clone, and transfected into HeLa cells. The hypothetical CVB5-P1anc assembled into functional virus particles that displayed phenotypic properties similar to those of contemporary clinical isolates. This is the first report describing the reconstruction and characterization of a fully functional picornavirus with a putative ancestral capsid.  相似文献   

11.
The combinatorial nature of genetic recombination can potentially provide organisms with immediate access to many more positions in sequence space than can be reached by mutation alone. Recombination features particularly prominently in the evolution of a diverse range of viruses. Despite rapid progress having been made in the characterization of discrete recombination events for many species, little is currently known about either gross patterns of recombination across related virus families or the underlying processes that determine genome-wide recombination breakpoint distributions observable in nature. It has been hypothesized that the networks of coevolved molecular interactions that define the epistatic architectures of virus genomes might be damaged by recombination and therefore that selection strongly influences observable recombination patterns. For recombinants to thrive in nature, it is probably important that the portions of their genomes that they have inherited from different parents work well together. Here we describe a comparative analysis of recombination breakpoint distributions within the genomes of diverse single-stranded DNA (ssDNA) virus families. We show that whereas nonrandom breakpoint distributions in ssDNA virus genomes are partially attributable to mechanistic aspects of the recombination process, there is also a significant tendency for recombination breakpoints to fall either outside or on the peripheries of genes. In particular, we found significantly fewer recombination breakpoints within structural protein genes than within other gene types. Collectively, these results imply that natural selection acting against viruses expressing recombinant proteins is a major determinant of nonrandom recombination breakpoint distributions observable in most ssDNA virus families.Genetic recombination is a ubiquitous biological process that is both central to DNA repair pathways (10, 57) and an important evolutionary mechanism. By generating novel combinations of preexisting nucleotide polymorphisms, recombination can potentially accelerate evolution by increasing the population-wide genetic diversity upon which adaptive selection relies. Recombination can paradoxically also prevent the progressive accumulation of harmful mutations within individual genomes (18, 35, 53). Whereas its ability to defend high-fitness genomes from mutational decay possibly underlies the evolutionary value of sexuality in higher organisms, in many microbial species where pseudosexual genetic exchange is permissible among even highly divergent genomes, recombination can enable access to evolutionary innovations that would otherwise be inaccessible by mutation alone.Such interspecies recombination is fairly common in many virus families (8, 17, 27, 44, 82). It is becoming clear, however, that as with mutation events, most recombination events between distantly related genomes are maladaptive (5, 13, 38, 50, 63, 80). As genetic distances between parental genomes increase, so too does the probability of fitness defects in their recombinant offspring (16, 51). The viability of recombinants is apparently largely dependent on how severely recombination disrupts coevolved intragenome interaction networks (16, 32, 51). These networks include interacting nucleotide sequences that form secondary structures, sequence-specific protein-DNA interactions, interprotein interactions, and amino acid-amino acid interactions within protein three-dimensional folds.One virus family where such interaction networks appear to have a large impact on patterns of natural interspecies recombination are the single-stranded DNA (ssDNA) geminiviruses. As with other ssDNA viruses, recombination is very common among the species of this family (62, 84). Partially conserved recombination hot and cold spots have been detected in different genera (39, 81) and are apparently caused by both differential mechanistic predispositions of genome regions to recombination and natural selection disfavoring the survival of recombinants with disrupted intragenome interaction networks (38, 51).Genome organization and rolling circle replication (RCR)—the mechanism by which geminiviruses and many other ssDNA viruses replicate (9, 67, 79; see reference 24 for a review)—seem to have a large influence on basal recombination rates in different parts of geminivirus genomes (20, 33, 39, 61, 81). To initiate RCR, virion-strand ssDNA molecules are converted by host-mediated pathways into double-stranded “replicative-form” (RF) DNAs (34, 67). Initiated by a virus-encoded replication-associated protein (Rep) at a well-defined virion-strand replication origin (v-ori), new virion strands are synthesized on the complementary strand of RF DNAs (28, 73, 74) by host DNA polymerases. Virion-strand replication is concomitant with the displacement of old virion strands, which, once complete, yields covalently closed ssDNA molecules which are either encapsidated or converted into additional RF DNAs. Genome-wide basal recombination rates in ssDNA viruses are probably strongly influenced by the specific characteristics of host DNA polymerases that enable RCR. Interruption of RCR has been implicated directly in geminivirus recombination (40) and is most likely responsible for increased basal recombination rates both within genes transcribed in the opposite direction from that of virion-strand replication (40, 71) and at the v-ori (1, 9, 20, 69, 74).Whereas most ssDNA virus families replicate via either a rolling circle mechanism (the Nanoviridae, Microviridae, and Geminiviridae) (3, 23, 24, 31, 59, 67, 74) or a related rolling hairpin mechanism (the Parvoviridae) (25, 76), among the Circoviridae only the Circovirus genus is known to use RCR (45). Although the Gyrovirus genus (the other member of the Circoviridae) and the anelloviruses (a currently unclassified ssDNA virus group) might also use RCR, it is currently unknown whether they do or not (78). Additionally, some members of the Begomovirus genus of the Geminiviridae either have a second genome component, called DNA-B, or are associated with satellite ssDNA molecules called DNA-1 and DNA-Beta, all of which also replicate by RCR (1, 47, 68).Recombination is known to occur in the parvoviruses (19, 43, 70), microviruses (66), anelloviruses (40, 46), circoviruses (11, 26, 60), nanoviruses (30), geminivirus DNA-B components, and geminivirus satellite molecules (2, 62). Given that most, if not all, of these ssDNA replicons are evolutionarily related to and share many biological features with the geminiviruses (22, 31, 36), it is of interest to determine whether conserved recombination patterns observed in the geminiviruses (61, 81) are evident in these other groups. To date, no comparative analyses have ever been performed with different ssDNA virus families to identify, for example, possible influences of genome organization on recombination breakpoint distributions found in these viruses.Here we compare recombination frequencies and recombination breakpoint distributions in most currently described ssDNA viruses and satellite molecules and identify a number of sequence exchange patterns that are broadly conserved across this entire group.  相似文献   

12.
The multiplicity of infection (MOI), i.e., the number of virus genomes that infect a cell, is a key parameter in virus evolution, as it determines processes such as genetic exchange among genomes, selection intensity on viral genes, epistatic interactions, and the evolution of multipartite viruses. In fact, the MOI level is equivalent to the virus ploidy during genome expression. Nevertheless, there are few experimental estimates of MOI, particularly for viruses with eukaryotic hosts. Here we estimate the MOI of Tobacco mosaic virus (TMV) in its systemic host, Nicotiana benthamiana. The progress of infection of two TMV genotypes, differently tagged with the green or red fluorescent proteins GFP and RFP, was monitored by determining the number of leaf cell protoplasts that showed GFP, RFP, or GFP and RFP fluorescence at different times postinoculation. This approach allowed the quantitative analysis of the kinetics of infection and estimation of the generation time and the number of infection cycles required for leaf colonization. MOI levels were estimated from the frequency of cells infected by only TMV-GFP or TMV-RFP. The MOI was high, but it changed during the infection process, decreasing from an initial level of about 6 to a final one of 1 to 2, with most infection cycles occurring at the higher MOI levels. The decreasing MOI can be explained by mechanisms limiting superinfection and/or by genotype competition within double-infected cells, which was shown to occur in coinfected tobacco protoplasts. To our knowledge, this is the first estimate of MOI during virus colonization of a eukaryotic host.Virus evolution has been a very active area of research in the last few decades, as viruses are both important pathogens of humans, animals, and plants and good models to experimentally test hypotheses on parasite evolution or, more generally, central questions on evolutionary biology (11, 12, 21, 36). Considerable efforts have been devoted to modeling the evolution of viral populations. However, contrasting the theoretical models with reality may be hindered by limited experimental information on important parameters of the virus life cycle. The multiplicity of infection (MOI), i.e., the number of virus particles or genomes that may infect a cell, is a key parameter in many models of virus evolution (5, 6, 14, 15, 37, 38, 39, 52, 53, 57, 61) for which experimental estimates are scant.When a cell is coinfected by different viral genomes, competition may lead to decreased fitness of individual genotypes in comparison with their fitness in single infections (15, 31, 40). Thus, limiting coinfection may result in a selective advantage for viruses (58), which have developed mechanisms to prevent superinfection of previously infected cells (51, 60). On the other hand, infection of a cell by more than one virus genome is a prerequisite for two central phenomena in virus genetics to take place: recombination and complementation of defective mutants. Recombination between viral strains during replication in the same cell and complementation of defective mutants have been extensively documented for viruses infecting prokaryotes, animals, and plants (2, 25, 56), indicating that there must be some degree of coinfection and, hence, that the MOI must be higher than one in at least some infected cells. However, estimates of MOI in the natural hosts of viruses are surprisingly scarce in spite of this parameter''s relevance: values of about 2 to 3 have been reported for different DNA or RNA bacteriophages (26, 41, 51, 58), and a value of 4 to 5 was reported for Autographa californica nuclear polyhedrosis virus infecting larvae of the moth Tricoplusia ni (3), to our knowledge, the only estimate for a virus in its eukaryotic host. We are not aware of estimates reported for viruses infecting mammals or plants, although a MOI of about 3 can be inferred from the number of proviral copies of HIV in spleen cells of infected patients (29). This paucity of data may be due to the technical difficulty of directly measuring MOI, particularly within a eukaryotic host. Genetic approaches may provide valid alternatives for estimating MOI levels (3, 58), and here, the MOI of a plant virus is estimated through the analysis of the relative frequencies of two genotypes during the process of host colonization.Host colonization by plant-infecting viruses has been known for a long time to be a two-step phenomenon. First, colonization proceeds slowly from the initially infected cells to their neighbors by way of the cytoplasmic connections called plasmodesmata, a process known as cell-to-cell movement. After infection thus reaches the cells in the vasculature, the second step, known as long-distance or systemic movement, occurs as viruses move faster to distant organs through the vascular tissue, the phloem in most cases (59). As a result of these processes, the virus population within the infected plant may be strongly structured. Analyses of different viruses in different host plant species have shown that systemic movement causes population bottlenecks that may be severe (16, 28, 32, 34, 46), resulting in differences in the genetic composition of the virus subpopulations in different systemically infected organs. No analysis of population bottlenecks during cell-to-cell movement has been reported, but data indicate that the virus population within a leaf has a strong spatial structure with a separate distribution of different genotypes in different leaf areas. These reports derive from analyses of viruses that differ in genomic organization and gene expression strategies in different host plant species (9, 10, 23, 55); they indicate that a separate distribution of viral genotypes within the infected leaf is a general phenomenon and suggest limitation of coinfection. Data on the spatial exclusion of virus genotypes within the infected leaf are in apparent contradiction with the abundant evidence of recombination and complementation of defective mutants, which has been widely documented for plant viruses (19, 44, 50, 62). It should be pointed out that all reports on the spatial exclusion of virus genotypes in an infected leaf derive from microscopy observations, mostly at late times after infection of the tissue. No information is available on the kinetics of leaf colonization by viruses, and current data do not allow the estimation of MOI.In this report, we estimate the MOI of a plant RNA virus, Tobacco mosaic virus (TMV), in its systemic host, Nicotiana benthamiana. For this, we have reexamined the process of virus colonization by monitoring the progress of infection of two TMV genotypes in inoculated and in systemically infected leaves. The two TMV genotypes differed in the expression of fluorescent tags, either the green fluorescent protein (GFP) from Aequorea victoria (42, 43) or a red fluorescent protein (RFP) from Discosoma sp. (49). The expression of GFP and RFP allowed the precise quantification of the number of cells infected by either one or both TMV genotypes, and these data allowed the estimation of genotype frequencies and of MOI. The results show evidence of strong spatial structure of the virus population, with most cells being infected by either TMV-GFP or TMV-RFP alone and only a small fraction of cells being double infected. The kinetics of the single and double infections show that the MOI changes with time, decreasing as colonization progresses and therefore suggesting that exclusion mechanisms operate at later times after infection.  相似文献   

13.
Poxviruses produce complement regulatory proteins to subvert the host''s immune response. Similar to the human pathogen variola virus, ectromelia virus has a limited host range and provides a mouse model where the virus and the host''s immune response have coevolved. We previously demonstrated that multiple components (C3, C4, and factor B) of the classical and alternative pathways are required to survive ectromelia virus infection. Complement''s role in the innate and adaptive immune responses likely drove the evolution of a virus-encoded virulence factor that regulates complement activation. In this study, we characterized the ectromelia virus inhibitor of complement enzymes (EMICE). Recombinant EMICE regulated complement activation on the surface of CHO cells, and it protected complement-sensitive intracellular mature virions (IMV) from neutralization in vitro. It accomplished this by serving as a cofactor for the inactivation of C3b and C4b and by dissociating the catalytic domain of the classical pathway C3 convertase. Infected murine cells initiated synthesis of EMICE within 4 to 6 h postinoculation. The levels were sufficient in the supernatant to protect the IMV, upon release, from complement-mediated neutralization. EMICE on the surface of infected murine cells also reduced complement activation by the alternative pathway. In contrast, classical pathway activation by high-titer antibody overwhelmed EMICE''s regulatory capacity. These results suggest that EMICE''s role is early during infection when it counteracts the innate immune response. In summary, ectromelia virus produced EMICE within a few hours of an infection, and EMICE in turn decreased complement activation on IMV and infected cells.Poxviruses encode in their large double-stranded DNA genomes many factors that modify the immune system (30, 56). The analysis of these molecules has revealed a delicate balance between viral pathogenesis and the host''s immune response (2, 21, 31, 61). Variola, vaccinia, monkeypox, cowpox, and ectromelia (ECTV) viruses each produce an orthologous complement regulatory protein (poxviral inhibitor of complement enzymes [PICE]) that has structural and functional homology to host proteins (14, 29, 34, 38, 41, 45, 54). The loss of the regulatory protein resulted in smaller local lesions with vaccinia virus lacking the vaccinia virus complement control protein (VCP) (29) and in a greater local inflammatory response in the case of cowpox lacking the inflammation-modulatory protein (IMP; the cowpox virus PICE) (35, 45, 46). Additionally, the complete loss of the monkeypox virus inhibitor of complement enzymes (MOPICE) may account for part of the reduced mortality observed in the West African compared to Congo basin strains of monkeypox virus (12).The complement system consists of proteins on the cell surface and in blood that recognize and destroy invading pathogens and infected host cells (36, 52). Viruses protect themselves from the antiviral effects of complement activation in a variety of ways, including hijacking the host''s complement regulatory proteins or producing their own inhibitors (7, 8, 15, 20, 23). Another effective strategy is to incorporate the host''s complement regulators in the outermost viral membrane, which then protects the virus from complement attack (62). The extracellular enveloped virus (EEV) produced by poxviruses acquires a unique outer membrane derived from the Golgi complex or early endosomes that contain the protective host complement regulators (58, 62). Poxviruses have multiple infectious forms, and the most abundant, intracellular mature virions (IMV), are released when infected cells lyse (58). The IMV lacks the outermost membrane found on EEV and is sensitive to complement-mediated neutralization. The multiple strategies viruses have evolved to evade the complement system underscore its importance to innate and adaptive immunity (15, 36).The most well-characterized PICE is VCP (24-29, 34, 49, 50, 53, 55, 59, 60). Originally described as a secreted complement inhibitor (34), VCP also attaches to the surface of infected cells through an interaction with the viral membrane protein A56 that requires an unpaired N-terminal cysteine (26). This extra cysteine also adds to the potency of the inhibitor by forming function-enhancing dimers (41). VCP and the smallpox virus inhibitor of complement enzymes (SPICE) bind heparin in vitro, and this may facilitate cell surface interactions (24, 38, 50, 59). The coevolution of variola virus with its only natural host, humans, likely explains the enhanced activity against human complement observed with SPICE compared to the other PICEs (54, 64).Our recent work with ECTV, the causative agent of mousepox infection, demonstrated that the classical and alternative pathways of the complement system are required for host survival (48). The mouse-specific pathogen ECTV causes severe disease in most strains and has coevolved with its natural host, analogous to variola virus in humans (9). This close host-virus relationship is particularly important for evaluating the role of the complement system, given the species specificity of many complement proteins, receptors, and regulators (10, 47, 62). Additionally, the availability of complement-deficient mice permits dissection of the complement activation pathways involved. Naïve C57BL/6 mouse serum neutralizes the IMV of ECTV in vitro, predominately through opsonization (48). Maximal neutralization requires natural antibody, classical-pathway activation, and amplification by the alternative pathway. C3 deficiency in the normally resistant C57BL/6 strain results in acute mortality, similar to immunodeficiencies in important elements of the antiviral immune response, including CD8+ T cells (19, 32), natural killer cells (18, 51), and gamma interferon (33). During ECTV infection, the complement system acts in the first few hours and days to delay the spread of infection, resulting in lower levels of viremia and viral burden in tissues (48).This study characterized the PICE produced by ECTV, ectromelia virus inhibitor of complement enzymes (EMICE), and assessed its complement regulatory activity. Recombinant EMICE (rEMICE) decreased activation of both human and mouse complement. Murine cells produced EMICE at 4 to 6 h postinfection prior to the release of the majority of the complement-sensitive IMV from infected cells. rEMICE protected ECTV IMV from complement-mediated neutralization. Further, EMICE produced during natural infection inhibited complement deposition on infected cells by the alternative pathway. ECTV likely produces this abundance of EMICE to protect both the IMV and infected cells.  相似文献   

14.
15.
We have sequenced the genome and identified the structural proteins and lipids of the novel membrane-containing, icosahedral virus P23-77 of Thermus thermophilus. P23-77 has an ∼17-kb circular double-stranded DNA genome, which was annotated to contain 37 putative genes. Virions were subjected to dissociation analysis, and five protein species were shown to associate with the internal viral membrane, while three were constituents of the protein capsid. Analysis of the bacteriophage genome revealed it to be evolutionarily related to another Thermus phage (IN93), archaeal Halobacterium plasmid (pHH205), a genetic element integrated into Haloarcula genome (designated here as IHP for integrated Haloarcula provirus), and the Haloarcula virus SH1. These genetic elements share two major capsid proteins and a putative packaging ATPase. The ATPase is similar with the ATPases found in the PRD1-type viruses, thus providing an evolutionary link to these viruses and furthering our knowledge on the origin of viruses.Three-dimensional structures of the major capsid proteins, as well as the architecture of the virion and the sequence similarity of putative genome packaging ATPases, have revealed unexpected evolutionary connection between virus families. Viruses infecting hosts residing in different domains of life (Bacteria, Archaea, and Eukarya) share common structural elements and possibly also ways to package the viral genome (8, 13, 41). It has been proposed that the set of genes responsible for virion assembly is a hallmark of the virus and is designated as the innate viral “self,” which may retain its identity through evolutionary times (5). Based on this, it is proposed that viruses can be classified into lineages that span the different domains of life. Therefore, the studies of new virus isolates might provide insights into the events that led to the origin of viruses and maybe even the origin of life itself (34, 40). However, viruses are known to be genetic mosaics (28), and these structural lineages therefore do not reflect the evolutionary history of all genes in a given virus. For example, the genome replication strategies vary significantly even in the currently established lineages (41) and, consequently, a structural approach does not point out to a specific form of replication in the ancestor. Nevertheless, as the proposal for a viral self is driven from information on viral structures and pathways of genome encapsidation, the ancestral form of the self was likely to be composed of a protective coat and the necessary mechanisms to incorporate the genetic material within the coat.Viruses structurally related to bacteriophage PRD1, a phage infecting gram-negative bacteria, have been identified in all three domains of life, and the lineage hypothesis was first proposed based on structural information on such viruses. Initially, PRD1 and human adenovirus were proposed to originate from a common ancestor mainly due to the same capsid organization (T=25) and the major coat protein topology, the trimeric double β-barrel fold (12). In addition, these viruses share a common vertex organization and replication mechanism (20, 31, 53, 63). PRD1 is an icosahedral virus with an inner membrane, whereas adenovirus lacks the membrane. Later, many viruses with similar double β-barrel fold in the major coat protein have been discovered and included to this viral lineage. For example, the fold is present in Paramecium bursaria Chlorella virus 1 (56) of algae, Bam35 (45) of gram-positive bacteria, PM2 (2) of gram-negative marine bacteria, and Sulfolobus turreted icosahedral virus (STIV) (38) of an archaeal host. Moreover, genomic analyses have revealed a common set of genes in a number of nucleocytoplasmic large DNA viruses. Chilo iridescent virus and African swine fever virus 1 are related to Paramecium bursaria Chlorella virus 1 and most probably share structural similarity to PRD1-type viruses (13, 30, 31, 68). The largest known viruses, represented by mimivirus and poxvirus, may also belong to this lineage (29, 77). Two euryarchaeal proviruses, TKV4 and MVV, are also proposed to belong to this lineage based on bioinformatic searches (42). The proposed PRD1-related viruses share the same basic architectural principles despite major differences in the host organisms and particle and genome sizes (1, 2, 38, 56). PM2, for example, has a genome of only 10 kbp, whereas mimivirus (infecting Acanthamoeba polyphaga) double-stranded DNA (dsDNA) genome is 1.2 Mbp in size (59).How many virion structure-based lineages might there be? This obviously relates to the number of protein folds that have the properties needed to make viral capsids. It has been noted that, in addition to PRD1-type viruses, at least tailed bacterial and archaeal viruses, as well as herpesviruses, share the same coat protein fold. Also, certain dsRNA viruses seem to have structural and functional similarities, although their hosts include bacteria and yeasts, as well as plants and animals (6, 18, 19, 27, 55, 60, 74). Obviously, many structural principles to build a virus capsid exist, and it has been suggested that especially geothermally heated environments have preserved many of the anciently formed virus morphotypes (35).Thermophilic dsDNA bacteriophage P23-77 was isolated from an alkaline hot spring in New Zealand on Thermus thermophilus (17) ATCC 33923 (deposited as Thermus flavus). P23-77 was shown to have an icosahedral capsid and possibly an internal membrane but no tail (81). Previously, another Thermus virus, IN93, with a similar morphology has been described (50). IN93 was inducible from a lysogenic strain of Thermus aquaticus TZ2, which was isolated from hot spring soil in Japan. Recently, P23-77 was characterized in more detail (33). It has an icosahedral protein coat, organized in a T=28 capsid lattice (21). The presence of an internal membrane was confirmed, and lipids were shown to be constituents of the virion. Ten structural proteins were identified, with apparent molecular masses ranging from 8 to 35 kDa. Two major protein species with molecular masses of 20 and 35 kDa were proposed to make the capsomers, one forming the hexagonal building blocks and the other the two towers that decorate the capsomer bases (33). Surprisingly, P23-77 is structurally closest to the haloarchaeal virus SH1, which is the only other example of a T=28 virion architecture (32, 33). In both cases it was proposed that the capsomers are made of six single β-barrels opposing the situation with the other structurally related viruses where the hexagonal capsomers are made of three double β-barrel coat protein monomers (8).In the present study we analyze the dsDNA genome of P23-77. Viral membrane proteins and those associated with the capsid were identified by virion dissociation studies. The protein chemistry data and genome annotation are consistent with the results of the disruption studies. A detailed analysis of the lipid composition of P23-77 and its T. thermophilus host was carried out. The data collected here reveal additional challenges in attempts to generate viral lineages based on the structural and architectural properties of the virion.  相似文献   

16.
17.
Only a few archaeal viruses have been subjected to detailed structural analyses. Major obstacles have been the extreme conditions such as high salinity or temperature needed for the propagation of these viruses. In addition, unusual morphotypes of many archaeal viruses have made it difficult to obtain further information on virion architectures. We used controlled virion dissociation to reveal the structural organization of Halorubrum pleomorphic virus 1 (HRPV-1) infecting an extremely halophilic archaeal host. The single-stranded DNA genome is enclosed in a pleomorphic membrane vesicle without detected nucleoproteins. VP4, the larger major structural protein of HRPV-1, forms glycosylated spikes on the virion surface and VP3, the smaller major structural protein, resides on the inner surface of the membrane vesicle. Together, these proteins organize the structure of the membrane vesicle. Quantitative lipid comparison of HRPV-1 and its host Halorubrum sp. revealed that HRPV-1 acquires lipids nonselectively from the host cell membrane, which is typical of pleomorphic enveloped viruses.In recent years there has been growing interest in viruses infecting hosts in the domain Archaea (43). Archaeal viruses were discovered 35 years ago (52), and today about 50 such viruses are known (43). They represent highly diverse virion morphotypes in contrast to the vast majority (96%) of head-tail virions among the over 5,000 described bacterial viruses (1). Although archaea are widespread in both moderate and extreme environments (13), viruses have been isolated only for halophiles and anaerobic methanogenes of the kingdom Euryarchaeota and hyperthermophiles of the kingdom Crenarchaeota (43).In addition to soil and marine environments, high viral abundance has also been detected in hypersaline habitats such as salterns (i.e., a multipond system where seawater is evaporated for the production of salt) (19, 37, 50). Archaea are dominant organisms at extreme salinities (36), and about 20 haloarchaeal viruses have been isolated to date (43). The majority of these are head-tail viruses, whereas electron microscopic (EM) studies of highly saline environments indicate that the two other described morphotypes, spindle-shaped and round particles, are the most abundant ones (19, 37, 43). Thus far, the morphological diversity of the isolated haloarchaeal viruses is restricted compared to viruses infecting hyperthermophilic archaea, which are classified into seven viral families (43).All of the previously described archaeal viruses have a double-stranded DNA (dsDNA) genome (44). However, a newly characterized haloarchaeal virus, Halorubrum pleomorphic virus 1 (HRPV-1), has a single-stranded DNA (ssDNA) genome (39). HRPV-1 and its host Halorubrum sp. were isolated from an Italian (Trapani, Sicily) solar saltern. Most of the studied haloarchaeal viruses lyse their host cells, but persistent infections are also typical (40, 44). HRPV-1 is a nonlytic virus that persists in the host cells. In liquid propagation, nonsynchronous infection cycles of HRPV-1 lead to continuous virus production until the growth of the host ceases, resulting in high virus titers in the growth medium (39).The pleomorphic virion of HRPV-1 represents a novel archaeal virus morphotype constituted of lipids and two major structural proteins VP3 (11 kDa) and VP4 (65 kDa). The genome of HRPV-1 is a circular ssDNA molecule (7,048 nucleotides [nt]) containing nine putative open reading frames (ORFs). Three of them are confirmed to encode structural proteins VP3, VP4, and VP8, which is a putative ATPase (39). The ORFs of the HRPV-1 genome show significant similarity, at the amino acid level, to the minimal replicon of plasmid pHK2 of Haloferax sp. (20, 39). Furthermore, an ∼4-kb region, encoding VP4- and VP8-like proteins, is found in the genomes of two haloarchaea, Haloarcula marismortui and Natronomonas pharaonis, and in the linear dsDNA genome (16 kb) of spindle-shaped haloarchaeal virus His2 (39). The possible relationship between ssDNA virus HRPV-1 and dsDNA virus His2 challenges the classification of viruses, which is based on the genome type among other criteria (15, 39).HRPV-1 is proposed to represent a new lineage of pleomorphic enveloped viruses (39). A putative representative of this lineage among bacterial viruses might be L172 of Acholeplasma laidlawii (14). The enveloped virion of L172 is pleomorphic, and the virus has a circular ssDNA genome (14 kb). In addition, the structural protein pattern of L172 with two major structural proteins, of 15 and 53 kDa, resembles that of HRPV-1.The structural approach has made it possible to reveal relationships between viruses where no sequence similarity can be detected. It has been realized that several icosahedral viruses infecting hosts in different domains of life share common virion architectures and folds of their major capsid proteins. These findings have consequences for the concept of the origin of viruses. A viral lineage hypothesis predicts that viruses within the same lineage may have a common ancestor that existed before the separation of the cellular domains of life (3, 5, 8, 26). Currently, limited information is available on the detailed structures of viruses infecting archaea. For example, the virion structures of nontailed icosahedral Sulfolobus turreted icosahedral virus (STIV) and SH1 have been determined (21, 23, 46). However, most archaeal viruses represent unusual, sometimes nonregular, morphotypes (43), which makes it difficult to apply structural methods that are based on averaging techniques.A biochemical approach, i.e., controlled virion dissociation, gives information on the localization and interaction of virion components. In the present study, controlled dissociation was used to address the virion architecture of HRPV-1. A comparative lipid analysis of HRPV-1 and its host was also carried out. Our results show that the unique virion type is composed of a flexible membrane decorated with the glycosylated spikes of VP4 and internal membrane protein VP3. The circular ssDNA genome resides inside the viral membrane vesicle without detected association to any nucleoproteins.  相似文献   

18.
19.
20.
There is an urgent need for human immunodeficiency virus (HIV) vaccines that induce robust mucosal immunity. Influenza A viruses (both H1N1 and H3N2) were engineered to express simian immunodeficiency virus (SIV) CD8 T-cell epitopes and evaluated following administration to the respiratory tracts of 11 pigtail macaques. Influenza virus was readily detected from respiratory tract secretions, although the infections were asymptomatic. Animals seroconverted to influenza virus and generated CD8 and CD4 T-cell responses to influenza virus proteins. SIV-specific CD8 T-cell responses bearing the mucosal homing marker β7 integrin were induced by vaccination of naïve animals. Further, SIV-specific CD8 T-cell responses could be boosted by recombinant influenza virus-SIV vaccination of animals with already-established SIV infection. Sequential vaccination with influenza virus-SIV recombinants of different subtypes (H1N1 followed by H3N2 or vice versa) produced only a limited boost in immunity, probably reflecting T-cell immunity to conserved internal proteins of influenza A virus. SIV challenge of macaques vaccinated with an influenza virus expressing a single SIV CD8 T cell resulted in a large anamnestic recall CD8 T-cell response, but immune escape rapidly ensued and there was no impact on chronic SIV viremia. Although our results suggest that influenza virus-HIV vaccines hold promise for the induction of mucosal immunity to HIV, broader antigen cover will be needed to limit cytotoxic T-lymphocyte escape.Developing a safe and effective human immunodeficiency virus (HIV) vaccine is one of the defining scientific challenges of our time. Induction of peripheral CD8 T-cell immunity to HIV did not protect against sexual exposure to HIV type 1 (HIV-1) in humans in a recent efficacy trial (11, 43). In simian immunodeficiency virus (SIV)-macaque studies, peripheral CD8 T-cell immunity can effectively control viremia (40) but is often observed to have a transient or limited role in delaying SIV disease in macaques (32). The gradual accumulation of immune escape at CD8 T-cell epitopes undermines the effectiveness of CD8 T-cell immunity to SIV (6, 22, 46). It is likely that inducing mucosal CD8 T-cell immunity to HIV will be more effective at limiting viral replication during the very early phases of acute infection, prior to massive viral dissemination and destruction of large numbers of CD4 T cells (50). The induction of multifunctional mucosal CD8 T cells by live attenuated SIV vaccination of macaques is thought to play a significant role in the success of this strategy (25, 26); however, it is unfortunately too dangerous for clinical trials at present.A series of mucosal viral and bacterial HIV vaccine vectors have been studied in recent years; however, none have yet proceeded to advanced clinical trials. Live attenuated poliovirus vectors have shown promise in SIV studies, but these viruses can in rare cases revert to virulence (14). Salmonella-based SIV vaccine vectors are able to induce CD8 T-cell responses which express the α4β7 integrin mucosal homing marker when administered orally (20, 24). However, there may be a much stronger link between concomitant genital tract immunity and immunity induced at respiratory mucosal sites compared to that induced at enteric sites (33, 38, 42). Vesicular stomatitis virus vectors that replicate in the nasal mucosa show promise in SIV-macaque trials but are potentially neurotoxic (55). Replication-competent adenovirus vectors have looked promising in some SHIV-macaque studies (49) but failed to provide significant protection in a recent SIV-macaque study (17) and could have similar issues of enhanced infection rates as seen in the recent efficacy trials of replication-incompetent adenovirus type 5 vectors.A mucosal vector system that has several advantages over existing models but that is relatively unexplored is recombinant attenuated influenza viruses. Such viruses (i) have an existing reverse genetics system to readily generate and manipulate recombinant viruses (31, 34), (ii) are effective as anti-influenza vaccines and licensed for human use (e.g., “Flumist” vaccine [9]) with ready production capability, (iii) have robust respiratory mucosal replication that should facilitate genital mucosal immunity, and (iv) can be generated with a variety of hemagglutinin (H) and neuraminidase (N) glycoproteins, potentially enabling these viruses to be administered sequentially in prime-boost combinations to limit the effect of antivector humoral immunity (34). Mouse-adapted recombinant influenza virus-HIV vectors have been studied in mice and demonstrated significant induction of cellular immunity at mucosal sites (8, 27, 28, 44, 48). However, although several native influenza viruses replicate efficiently in the respiratory tracts of Asian macaque species (10, 12, 52), no studies to date have examined the immunogenicity or efficacy of recombinant attenuated influenza virus-SIV vectors in macaques.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号